Sie sind auf Seite 1von 9

Chemosphere 161 (2016) 349e357

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Modeling the oxidation of phenolic compounds by hydrogen peroxide


photolysis
ez*
Tianqi Zhang, Long Cheng, Lin Ma, Fanchao Meng, Robert G. Arnold, A. Eduardo Sa
Department of Chemical and Environmental Engineering, University of Arizona, Tucson, AZ, 85721, United States

h i g h l i g h t s

 A mechanism for oxidation of phenolic compounds by hydroxyl radicals is presented.


 Hydroxyl radical scavenging by intermediates reduces effectiveness of process.
 Kinetic model of the full mechanism is necessary for prediction of reactor performance.

a r t i c l e i n f o a b s t r a c t

Article history: Hydrogen peroxide UV photolysis is among the most widely used advanced oxidation processes (AOPs)
Received 7 March 2016 for the destruction of trace organics in waters destined for reuse. Previous kinetic models of hydrogen
Received in revised form peroxide photolysis focus on the dynamics of hydroxyl radical production and consumption, as well as
13 June 2016
the reaction of the target organic with hydroxyl radicals. However, the rate of target destruction may also
Accepted 29 June 2016
be affected by radical scavenging by reaction products. In this work, we build a predictive kinetic model
for the destruction of p-cresol by hydrogen peroxide photolysis based on a complete reaction mechanism
Handling Editor: I. Cousins that includes reactions of intermediates with hydroxyl radicals. The results show that development of a
predictive kinetic model to evaluate process performance requires consideration of the complete reac-
Keywords: tion mechanism, including reactions of intermediates with hydroxyl radicals.
Advanced oxidation processes 2016 Elsevier Ltd. All rights reserved.
Kinetic model
Phenolic compounds
Radical reactions

1. Introduction treatment and following the release of treated wastewater into the
environment is of practical importance.
Alkylphenol ethoxylates are non-ionic surfactants commonly Advanced oxidation processes (AOPs) are effective treatments
used to formulate products such as detergents, paints, dispersing for phenolic compounds due to their high rates of reaction with
agents, wetting products, pesticides, and lubricants (Nagarnaik and hydroxyl radicals (second-order rate constants in the range
Boulanger, 2011). These compounds are converted into alkylphe- 109e1010 M1 s1; Buxton et al., 1988). The UV photolysis of
nols in bioreactors used in wastewater treatment (Ahel et al., 1996), hydrogen peroxide (UV/H2O2) is among the most widely used AOPs
leading to their presence in surface waters (Scullion et al., 1996; in water treatment. Various models have been developed to
Dong et al., 2015), groundwater (Swartz et al., 2006), municipal simulate process kinetics. Common elements in these models
and industrial wastewater (Loyo-Rosales et al., 2007), aquatic include: (i) the kinetics of radical production from the direct
sediments and marine shellsh (Lye et al., 1999). The most photolysis of hydrogen peroxide, and hydroxyl radical scavenging
commonly detected metabolite, p-nonylphenol, induces breast tu- by hydrogen peroxide and other radicals; (ii) reaction of hydroxyl
mor cell proliferation (Soto et al., 1991) and is a recognized endo- radicals with the target compound and, although to a lesser extent,
crine disrupter (Lee and Lee, 1996). Therefore, the fates of p- (iii) radical scavenging by reaction byproducts and intermediates
nonylphenol and other phenolic compounds during wastewater (Lay, 1989; Glaze et al., 1995; Liao and Gurol, 1995; Stefan et al.,
1996; Crittenden et al., 1999; Murcia et al., 2015; Rojas et al.,
2010, 2011).
* Corresponding author. The rst step in the reaction mechanism for radical oxidation of
ez).
E-mail address: esaez@email.arizona.edu (A.E. Sa phenolic compounds is OH addition to the ring (Wojn arovits et al.,

http://dx.doi.org/10.1016/j.chemosphere.2016.06.110
0045-6535/ 2016 Elsevier Ltd. All rights reserved.
350 T. Zhang et al. / Chemosphere 161 (2016) 349e357

2002). The position of OH addition is controlled by substitutions on methylcatechol (Acros, 98%), hydrogen peroxide (Acros, 50 wt%),
the ring, but para- and ortho-hydroxylation normally predominate titanium sulfate (Pfaltz&Bauer, 30%), formate (Fluka, 1000 mg/
due to the electrophilic nature of the hydroxyl radical (Omura and L 4 mg/L), acetate (Fluka, 1000 mg/L 4 mg/L), oxalate (Fluka,
Matsuura, 1968; Matsuura and Omura, 1974). Further oxidation of 1000 mg/L 4 mg/L), and TOC standard (Fluka, 50 mg/L 2 mg/L).
the resulting dihydroxy compound occurs via abstraction of a Methanol (Fisher Chemical, HPLC Grade) and acetonitrile (Fisher
hydrogen atom to form benzoquinones, after which the ring can be Chemical, HPLC Grade) were solvents for HPLC analyses. Milli-Q
cleaved to yield aliphatic acids (Scheck and Frimmel, 1995). water (resistivity 18.0 mU cm1) was obtained from a Barnstead
Previous efforts on the kinetic modeling of the UV/H2O2 process NANOpure II system. Glassware was washed and heated overnight
generally include a complete description of the chemistry behind at 550  C prior to use.
radical formation, but rarely consider reaction intermediate and by-
products. Glaze et al. (1995) proposed a kinetic model containing 2.2. Experimental procedure
the most important elements of UV/H2O2 photolysis to predict the
time-dependent concentration of a halogenated target. They used a All experiments were conducted in a cylindrical (14-cm diam-
quasi-steady-state assumption (QSSA) to calculate the time- eter, 950-mL) glass batch reactor, equipped with low pressure (LP)
dependent concentrations of free radicals. Their model did not UV lamp, UV uence meter (Hamamatsu, H8025-254) and mag-
account for CO2 evolution/pH change and neglected the role of all netic stirrer. Stock solutions of p-cresol were prepared one day
radical scavenging species except H2O2. Stefan et al. (1996) fol- before use. 4-Methylcatechol stock solutions were prepared a half
lowed a similar approach, but assumed that the target compound hour before use in amber glass containers. The UV lamp (43-cm
(in their case acetone) was transformed to oxalic and formic acids long) was turned on two hours before the experiments. The lamp
and then to CO2. Subsequently, Crittenden et al. (1999) modeled position was 21 cm above the liquid level in the reactor. A quartz
UV/H2O2 process without reliance on a quasi-steady-state sleeve covered the lamp to cut-off wavelengths below 185 nm.
assumption for radical concentrations. Their work showed that Solution volume was 300 mL in all experiments (liquid depth was
the QSSA tends to under-predict hydroxyl radical concentrations in 1.95 cm). Experiments were initiated by exposing the reactor to UV
some cases and therefore the rates of target compound decompo- light. Experiment duration depended on the initial p-cresol con-
sition. Furthermore, their model included radical scavenging centration but ranged from one to three hours. Samples were
reactions. analyzed immediately after withdrawal. Dilution was frequently
The model of Song et al. (2008) accounted for time-dependent necessary to maintain concentrations within the recommended
radical concentrations, pH changes due to product formation (CO2 analytical range. Temperature was maintained at T 24.5 0.5  C.
and low molecular weight acids), and radical scavenging by H2O2 The trajectory of solution pH was determined in each experiment.
and natural organic matter.
More recent modeling efforts of oxidation of phenolic com- 2.3. Analytics
pounds usually include a full formulation of the dynamics of radical
formation, but very few or no details of the inuence of interme- The concentration of p-cresol was measured by uorescence
diate compounds and by-products over target compound degra- spectrometry (PerkinElmer LS 55). The inner lter effect (IFE)
dation. De Luis et al. (2011) presented a model of phenol and cresol caused by light absorption and reabsorption of light-absorbing
degradation during the UV/H2O2 process that considered possible compounds was corrected mathematically, accounting for the ab-
reactions of the organic compounds with produced radicals, but sorption spectrum of the sample (MacDonald et al., 1997). Solution
without considering intermediate compounds and by-products. A spectra were measured using a UVeVis spectrophotometer
similar approach was followed by Rojas et al. (2010, 2011) to model (Thermo Science, Genesys 10s). Even though the reaction inter-
the UV/H2O2 and Fentons reaction processes. Moreira et al. (2012) mediate 4-methylcatechol is produced in concentrations that are
developed a phenomenological model to simulate the TiO2 pho- comparable in order of magnitude to p-cresol, independent mea-
tocatalysis of phenol, which includes specic details about the re- surements showed that its uorescence intensities were negligible
action mechanism. compared to those of p-cresol in the relevant range of wavelengths.
In this work, p-cresol was selected as an example of a Analysis of primary intermediates, particularly 4-methyl-
substituted phenolic compound. Even though p-cresol itself is not a catechol, was carried out by HPLC (Agilent Technologies, 1200 Se-
common trace organic in wastewater efuent, it has chemical ries) with a Polar-RP, 80 A column (15 cm  4.6 mm  4 mm)
structure and reactivity with hydroxyl radicals similar to p-non- (Phenomenex) and UV/VIS detector at 280 nm. Elution solvents
ylphenol, which is a common endocrine disrupter compound were 10% methanol in water (A) and acetonitrile (B), Q 0.4 mL/
widely present in wastewater efuent. A mechanism for radical- min. The mobile phase initially consisted of 45% vol. B, linearly
dependent UV/photolysis was proposed and calibrated based on ramped to 65% vol. B at 15 min and then restored to the initial
time-dependent measurements of p-cresol and a number of reac- condition.
tion products/intermediates in a batch reactor. The model accounts Formate, acetate and oxalate were measured via ion chroma-
for changes in ionic strength and pH due to reactant conversions. tography (Dionex, ICS-5000) equipped with an AS18 column
Reactor uence rate and the p-cresol quantum yield during direct (2  250 mm) (Dionex IonPac), an AG18 guard column
photolysis were considered in the kinetic model. The calibrated (2  50 mm) (Dionex IonPac), and an AS-DV (Dionex) autosam-
model was employed to predict the evolution of concentrations of pler. Ions were detected by suppressed conductivity of the eluent
p-cresol and reaction intermediates under various experimental using an ASRS-2mm self-regenerating suppressor (Dionex). The
conditions. eluent (KOH solution; Q 0.4 mL/min) was programmed as fol-
lows: 1 mM for 8 min, ramped (linearly) to 30 mM in 20 min, then
2. Materials and methods immediately restored to 1 mM.
Hydrogen peroxide was measured using a modied perox-
2.1. Chemicals ytitanic (colorimetric) method (Boltz and Holwell, 1978). Absor-
bance was measured spectrophotometrically (Thermo Science,
All chemicals were obtained from commercial sources and used Genesys 10s) at 407 nm. Samples were diluted to maintain absor-
without further purication, including p-cresol (Acros, 99%), 4- bance values within linear limits. TOC was monitored using a Total
T. Zhang et al. / Chemosphere 161 (2016) 349e357 351

Organic Carbon analyzer (Shimadzu, TOC-VCSH). to carbon dioxide (Getoff et al., 1971; Leitner and Dore, 1997; Hart,
1951; Sun and Saeys, 2008).
2.4. Proposed mechanism
2.5. Kinetic model
The proposed mechanism of p-cresol oxidation by hydroxyl
radicals is shown in Fig. 1. The rst step is OH radical addition to the The proposed kinetic model for the oxidation of p-cresol during
ring to form 4-methylcatechol (Matsuura and Omura, 1974; Olariu UV/H2O2 photolysis consists of 36 reactions (Table 1). The UV/H2O2
et al., 2002). Two possible pathways (Omura and Matsuura, 1968; photolysis mechanism, including an initiation step that produces
Scheck and Frimmel, 1995) are used to explain the hydroxylation. hydroxyl radical (OH) and propagation steps that scavenge OH
One is the attack of OH radical on the phenolic ring, resulting in the and produce other active radicals (O 
2 and CO3 ), is based on the
formation of a cyclohexadienyl radical, which is subsequently work of Rojas et al. (2010, 2011) and references therein. Since the
converted into 4-methylcatechol. The other involves abstraction of reactor was open to the atmosphere, the initial carbon dioxide
a hydrogen atom yielding a phenoxyl radical that can combine with concentration was assumed to be at equilibrium with atmospheric
a second OH radical. However, the hydrogen abstraction reaction of carbon dioxide. The scavenging effect of inorganic carbon species,
a methyl H atom (Wang et al., 1998) is at least one order of bicarbonate and carbonate ions was considered in the model, but
magnitude slower than the formation of a cyclohexadienyl radical calculations showed that their effect was negligible in the range of
(Sehested et al., 1975; Wojna rovits et al., 2002). Further hydroxyl- conditions of the experiments.
ation may convert 4-methylcatechol; however, this reaction has The rate of direct photolysis of H2O2 is given by
negligible yield (Scheck and Frimmel, 1995). Oxidation of 4-
methylcatechol to 4-methylbenzoquinone occurs in two steps: RR1 FH2 O2 fH2 O2 I0 e2:303Az (1)
the OH radical attacks the ring forming a 4-methylcatechol semi-
quinone radical (Gohn and Getoff, 1977), after which the semi- where FH2 O2 (Table 1) is the quantum yield for the photocatalytic
quinone radical converts to 4-methylbenzoquinone in the presence decomposition of H2O2, fH2 O2 is the fraction of total light absorbed
of molecular oxygen (Davies et al., 1995). Once 4-methyl- by H2O2, I0 is the average uence rate (in Einsteins per unit time
benzoquinone is formed, ring tension and oxidizing agents force and reactor volume) at the reactor surface and Az is the total
ssion of the ring. The CeC bond in the ring between the carbons absorbance of the solution as function of depth. In the model, H2O2,
attached to oxygen atoms is cleaved, and 3-methyl-muconic acid is PC, 4 MC and 4 MB are the only important light-absorbing species
one of the ssion products (Pospisil et al., 1957; Scheck and at 254 nm. Therefore, Az is given by
Frimmel, 1995). Oxidation of 3-methyl-muconic acid involves the  
carbon double bonds. Because of the higher acidity of the a- Az H2 O2 H2 O2  PC PC 4MC 4MC 4MB 4MB z
hydrogen adjacent to the double bond, it can be easily abstracted by (2)
the OH radical (Jin et al., 2010; Huang et al., 2011). Low molecular
weight organic acids such as lactic and oxalic acid are oxidation where H2 O2 ; PC ; 4MC (Table 1) and 4MB (2100 M1 cm1; Waite,
products. The abstraction of a-hydrogen next to a eCOOH group by 1976; Jongberg et al., 2011) are molar extinction coefcients for
OH radical forms additional low-molecular -weight acids such as hydrogen peroxide, p-cresol, 4-methylcatechol and 4-methyl-
acetic and formic acids (Boonrattanakij et al., 2009). Eventually, benzoquinone at 254 nm, respectively, and z is the distance
oxalic acid, acetic acid and formic acid are oxidized by OH radicals measured downward within the reactive mixture from the top

Fig. 1. Proposed mechanism of p-cresol oxidation by hydroxyl radicals (see text for description).
352 T. Zhang et al. / Chemosphere 161 (2016) 349e357

Table 1
Kinetic and equilibrium parameters used in the model. Acronyms are dened in Fig. 1. Quantum yield (F) and molar extinction coefcients () are reported for photolysis
reactions, kinetic rate constants (k) are reported for relevant reactions, and equilibrium constants (K) are reported for reactions assumed to reach equilibrium instantaneously
in the reaction mixture.

Reaction Parameters Reference


1 1
R1 H2O2 hn / 2 OH 1 18.7 M cm Volman and Chen (1959)
F1 0.5 mol/Ein

R2 OH H2O2 / O 2 H2O H

k2 2.7  107 M1 s1 Buxton et al. (1988)

R3 OH HO 
2 / O2 H2O k3 7.5  109 M1 s1 Buxton et al. (1988)

R4 OH HCO 
3 / CO3 H2O k4 8.5  106 M1 s1 Buxton et al. (1988)
 
R5 OH CO2
3 / CO 3 OH k5 3.9  108 M1 s1 Buxton et al. (1988)

R6 OH HO2 / O2 H2O k6 6.6  109 M1 s1 Buxton et al. (1988)

R7 OH O
2 / O2 OH

k7 8.0  109 M1 s1 Buxton et al. (1988)

R8 OH OH / H2O2 k8 5.5  109 M1 s1 Buxton et al. (1988)

R9 OH CO
3 / products k9 3.0  109 M1 s1 Holeman et al. (1987)
R10 O
2 H2O2 / OH O2 OH

k10 0.13 M1 s1 Bielski et al. (1985)
 
R11 O2 CO3 / O2 CO2 3 k11 6.5  108 M1 s1 Eriksen et al. (1985)
R12 O 
2 HO2 / HO2 O2 k12 9.7  107 M1 s1 Bielski et al. (1985)
R13 HO2 HO2 / H2O2 O2 k13 8.6  105 M1 s1 Weinstein and Bielski (1979)
R14 HO2 H2O2 / OH O2 H2O k14 3.7 M1 s1 Bielski et al. (1985)
R15 CO 
3 H2O2 / HCO3 O2 H

k15 8.0  105 M1 s1 Neta et al. (1988)
R16 CO 
3 HO2 / HCO3 O2
 
k16 3.0  107 M1 s1 Neta et al. (1988)
R17 CO 
3 CO3 / 2CO3
2
k17 2.0  107 M1 s1 Neta et al. (1988)
R18 H2O2 4 HO 2 H

k18 2.51  102 s1 Perry et al. (1981)
k18r 1  1010 M1 s1
R19 HO2 4 O
2 H

k19 1.58  105 s1 Perry et al. (1981)
k19r 1  1010 M1 s1
R20 CO2(aq) H2O 4 H2CO3 k20 0.039 s1 Housecroft and Sharpe (2005)
k20r 23 s1
R21 H2CO3 4 HCO
3 H
k21 1.25  107 s1 Greenwood and Earnshaw (1997)
k21r 5  1010 M1 s1
R22 PC hn / products 22 340 M1 cm1 This work
F22 0.034 mol/Ein
R23 PC OH / 4 MC k23 1.2  1010 M1 s1 Buxton et al. (1988)
R24 4 MC hn / products 24 310 M1 cm1 This work
F24 0.007 mol/Ein
R25 4 MC OH / 4 MB k25 1.6  1010 M1 s1 Gohn and Getoff (1977)
R26 4 MB OH / 3 MM k26 2.0  1010 M1 s1 This work
R27 3 MM OH / L 2 Ox (Ox2) k27 5.0  108 M1 s1 Kang et al. (2002)
R28 L OH / HA (A) CO2 H2O k28 3.0  108 M1 s1 Buxton et al. (1988)
R29 Ox OH / Ox H2O k29 4.7  107 M1 s1 Buxton et al. (1988)
R30 Ox2 OH / Ox OH k30 7.7  106 M1 s1 Buxton et al. (1988)
R31 Ox Ox / Ox (Ox2) 2 CO2 k31 5.0  108 M1 s1 Ershov et al. (2008)
R32 HA OH / F CO2 H2O k32 1.23  108 M1 s1 Sun and Saeys (2008)
R33 A OH / F CO2 OH k33 8.5  107 M1 s1 Buxton et al. (1988)
R34 F OH / CO2 H2O k34 9.77  107 M1 s1 Sun and Saeys (2008)
E35 Ox 4 Ox2 H K35 6.31  105 M Benjamin (2002)
E36 HA 4 A H K36 1.74  105 M Benjamin (2002)
E37 HCO 2
3 4 CO3 H

K37 4.48  1011 M Benjamin (2002)

surface. Note that Acid-base reactions are considered to be at equilibrium. The


equilibrium equations are used to relate concentrations of the
H2 O2 zH2 O2  species involved. For example, for acetic acid,
fH2 O2 (3)
Az
The volume averaging of equation (1) over the batch reactor is,
after manipulations,  
TOTA g1 H
  HA   (6)
RR1 FH2 O2 fH2 O2 I0 1  e2:303A (4) K36 g1 H

where the angular brackets denote volume average, and


A Az z l, where l is the depth of the reacting mixture. Other
direct photolysis reactions are treated in the same manner.
TOTA K36
Radical reactions are generally second-order reactions. For g1 A    (7)
K36 g1 H
example, the reaction rate for the consumption of p-cresol by OH
radical is
where TOTA HA A , K36 is equilibrium constant of acetic acid
h i dissociation and g1 is the activity coefcient, which is calculated
_
RR23 k23 PC OH (5) from the Davies Equation.
Based on the mechanism proposed, the corresponding mole
where k23 is the second-order rate constant of p-cresol reaction balances of H2O2, p-cresol and 4-methylcatechol yield the following
with OH (Table 1). system of differential equations:
T. Zhang et al. / Chemosphere 161 (2016) 349e357 353

dH2 O2    h i
FH2 O2 fH2 O2 I0 1  e2:303A  k2 H2 O2  OH _
dt
h i2 h i h i2
k8 OH_  k10 H2 O2 g1 O_ _
2 k13 HO2
h i h i
 k14 H2 O2  HO_2  k15 H2 O2 g1 CO 3
_

  h i
 k18 H2 O2  k18r g1 HO 2 g1 H

(8)

dPC   h i
_
FPC fPC I0 1  e2:303A  k23 PC OH (9)
dt

d4MC   h i
_
F4MC f4MC I0 1  e2:303A k23 PC OH
dt
h i
 k 4MC OH _ (10)
25

TOC TOC0  CO2  CO2 0 (11)


Fig. 2. Decomposition of hydrogen peroxide in Milli-Q water in the batch reactor. Solid
where TOC0 and CO2 0 are initial TOC and CO2 molar concentra- lines: model predictions obtained by tting the model to decay curves using the u-
tions. More details of the model formulation, including mole bal- ence rate as adjustable parameter.
ances of radicals (OH, HO2, O 2 and CO
3 ) and reaction
intermediates (4 MB, 3 MM, LA , OA , OA2, OA, AA, AA and
 

FA), plus algebraic equations ([H2CO3] as a function of [TOTCO3] model to data representing compound decay. The solution absor-
and charge balance) are provided in the supporting information. bance for all experiments was measured as a function of time to
The fact that the reactor is well mixed implies that most species determine Az (equation (2)). This was necessary in direct photolysis
concentrations are uniform in the reactor. However, the relatively experiments since data showed that the products of direct
high reaction rates of radical species suggest that their concentra- photolysis reaction also absorb 254-nm light. (2,20 -Dihydroxy-5,5-
tions are not necessarily uniform. For example, the OH radical dimethylbiphenyl, 4-methylcatechol and 2-hydroxy-40 ,5-dime-
production by reaction R1 (Table 1) is not uniform due to the thyldiphenyl ether have been reported as products in the direct
attenuation of light with depth. In the model equations, radical photolysis of p-cresol, Joschek and Miller, 1966.) Results are pre-
concentrations are volume-averaged concentrations. The formula- sented in Fig. 3. To generate model predictions, we used the
tion is based on the assumption that kinetics that are not linear on calculated uence from the actinometry experiments (Fig. 2) and
radical concentrations (such as in radical/radical reactions like R8, tted the quantum yields to experimental data.
Table 1) can be expressed in terms of volume-averaged radical
concentrations. This approximation was explored in detail by Rojas
3.3. Model predictions: p-cresol degradation
et al. (2010).
The nal system of equations was solved using the stiff-
Fig. 4 provides predicted and experimental decay curves for p-
differential-equations solver ODE15s in MATLAB (based on Gears
cresol in a series of experiments carried out at different initial
method). The relative tolerance was 108, whereas the absolute
concentrations of H2O2 and p-cresol. Model calculations include the
tolerances for concentrations of all species were set four orders of
uence rate determined via actinometry and the quantum yields
magnitude below their respective values.
for p-cresol and 4-methylcatechol determined via direct photolysis.
In all cases experimental results are accurately predicted by the
3. Results and discussion model. Omitting OH radical scavenging by reaction intermediates
signicantly overestimates the destruction rate of p-cresol in most
3.1. Actinometry. Hydrogen peroxide photolysis cases (dashed lines in Fig. 4), which highlights the importance of
incorporating the full mechanism in the kinetic model.
The average uence rate at the surface of the reactor was Calculations show that intermediates, especially those before
determined using H2O2 actinometry. Hydrogen peroxide direct ring cleavage (4-methylcatechol and 4-methylbezoquinone) played
decomposition experiments were conducted at various initial an important role in hydroxyl radical quenching. There is no re-
concentrations. The kinetic model was applied to calculate the ported kinetic rate constant for the reaction of 4-methyl-
reactor surface average uence rate by tting experimental data for benzoquinone reaction with hydroxyl radical although Scheck and
H2O2 decay. Reactions 1e21 (Table 1) were used in the model Frimmel (1995) reported that 1,2-benzoquinone was not detected
formulation. A value of I0 0:85  106 Ein L1 s1 was obtained during oxidation of phenol by the UV/H2O2 process because it is
(Fig. 2), which was in agreement with the uence rate measured by easily cleaved into aliphatic compounds. In this study, the rate
the uence meter. constant for the reaction of 4-methylbenzoquinone with hydroxyl
radical was held to be of the same magnitude as that of 4-meth-
3.2. p-Cresol and 4-Methylcatechol direct photolysis ylcatechol, k26 2:0  1010 M 1 s1 , near the diffusion limit. Model
results were obtained with no additional tted parameters; all
Quantum yields of p-cresol and 4-methylcatechol were deter- other rate constants were obtained from preliminary experiments
mined from direct photolysis experiments for each compound (i.e. actinometry and direct photolysis experiments) or published
separately. In the experiments, p-cresol or 4-methylcatechol solu- data (Table 1).
tions were exposed to UV light in the absence of H2O2. The quan- Molecular oxygen participates in several steps of the reaction
tum yield of each compound was calculated by tting the kinetic mechanismdspecically, in the oxidation of 4-methylcatechol to 4-
354 T. Zhang et al. / Chemosphere 161 (2016) 349e357

Fig. 3. Decomposition of (a) p-cresol, and (b) 4-mehtylcatechol in aqueous solution by (LP)UV direct photolysis. Solid lines: model predictions obtained from a t of model results to
decay curves using the respective quantum yield as adjustable parameter. (a) [PC]0 0.438 mM, (b) [4 MC]0 0.103 mM.

Fig. 4. Degradation of p-cresol in aqueous solution by UV/H2O2 advanced oxidation as a function of initial H2O2 concentrations (a), and initial p-cresol concentrations (b). Solid
lines: full model predictions using the proposed mechanism; dashed lines: model calculations neglecting reactions of intermediates. (a) [PC]0 0.1 mM, (b) [H2O2]0 1 mM.

methylbenzoquinone (Davies et al., 1995) and the oxidation of experiments (Fig. 5). The reaction of p-cresol with OH radical was
carbon double bonds (Jin et al., 2010; Huang et al., 2011). To test the the only 4-methylcatechol source, but two sinks are considered in
potential role of oxygen, we performed experiments with degassed the kinetic model: direct photolysis and reaction with OH radical.
solutions by bubbling argon into the solution before exposure to UV Even though the rate constant for 4-methylcatechol reaction with
light. During the experiment, an argon atmosphere was maintained OH radical is higher than that of p-cresol, 4-methylcatechol accu-
over the reactor to prevent oxygen diffusion into the reacting mulates until most of the p-cresol is consumed. There was good
mixture. The results in terms of p-cresol concentration vs. time agreement between data and predictions at low H2O2 concentra-
showed no appreciable difference relative to the original experi- tions (Fig. 5a), but the model tends to underpredict concentrations
ments, suggesting that dissolved oxygen is not important in p- of 4-methylcatechol at later times and higher measured concen-
cresol oxidation. Representative results are provided in the sup- trations (Fig. 5b) while still trending correctly.
porting information. The evolutions of formic acid, acetic acid and oxalic acid con-
There are two possible pathways for the OH-induced oxidation centrations are presented in Fig. 6 for a representative experiment.
of p-cresol to 4-methylcatechol (Omura and Matsuura, 1968; (Results for p-cresol degradation in this experiment are in Fig. 5b.)
Scheck and Frimmel, 1995): the rst involves the cyclohexadienyl Although the a-hydrogens next to the carbon double bonds and
radical formed due to OH radical attack that converts the p-cresol carboxyl groups are more active under oxidizing conditions, there
into dihydroxy compounds. In the second, the phenoxyl radical may be several ways in which to break down the aliphatic acid
formed due to the abstraction of hydrogen atom subsequently formed by the ring cleavage. Therefore, the nal stages of the
combines with a second OH radical. However, the second mecha- mechanism proposed (Fig. 1) may lead to formation of other spe-
nism is dominant only when the OH-radical is present at high cies. Fig. 7 shows the trajectory of total carbon corresponding to
concentrations (Waters, 1964). In the kinetic model used here, only measured concentrations of aliphatic acids along with model cal-
one OH radical is involved in the conversion of p-cresol to 4- culations for a representative experiment. Both model and data
methylcatechol, suggesting that the cyclohexadienyl pathway indicate that aliphatic acids accumulate throughout the experi-
predominates. ment. By the end of the experiment (180 min), essentially all the p-
The production of 4-methylcatechol was measured in the cresol is destroyed (Fig. 5b).
T. Zhang et al. / Chemosphere 161 (2016) 349e357 355

Fig. 5. Degradation of p-cresol in aqueous solution by UV/H2O2 advanced oxidation. Evolution of 4-methylcatechol concentration with time. Solid lines: model predictions.
[PC]0 0.5 mM, (a) [H2O2]0 1 mM, (b) [H2O2]0 5 mM.

The concentrations of H2O2 and TOC, and pH were measured


throughout the experiment and predicted using the kinetic model
(Figs. 8e10). Despite some differences, the model adequately pre-
dicts trends in H2O2 concentration (Fig. 8) and pH (Fig. 9). Previous
studies have calculated pH changes during this type of experiment
by assuming that the target organics are completely converted into
carbon dioxide (Crittenden et al., 1999; Rojas et al., 2010). In our
study, a signicant drop in pH was observed in the 3-h experiment.
However, the drop of pH was due to the formation of organic acids.
The lack of CO2 production (that is, complete mineralization of the
target) is evident in the TOC results (Fig. 10) since TOC does not
change appreciably during the experiments, as predicted by our
model.

3.4. Model predictions: 4-Methylcatechol degradation

Experiments were carried out with 4-methylcatechol as the


Fig. 6. Evolution of formic acid, acetic acid and oxalic acid concentrations with time.
initial substrate at different initial hydrogen peroxide concentra-
[PC]0 0.5 mM, [H2O2]0 5 mM. tions. As anticipated, degradation rate was directly related to the
initial hydrogen peroxide concentration within the range of ex-
periments (Fig. 11). Agreement between experimental data and
model simulations is observed at 0.2 mM H2O2 but the model tends

Fig. 7. Evolution of aliphatic reaction intermediates concentration with time. Solid


line: model prediction. [PC]0 0.5 mM, [H2O2]0 5 mM. Abbreviations in Fig. 1.

Fig. 8. Evolution of H2O2 concentration with time. Solid line: model prediction.
[PC]0 0.5 mM, [H2O2]0 5 mM.
356 T. Zhang et al. / Chemosphere 161 (2016) 349e357

Fig. 11. Degradation of 4-methylcatechol in aqueous solution by UV/H2O2 advanced


Fig. 9. Evolution of pH with time. Solid line: model prediction. [PC]0 0.5 mM, oxidation as a function of initial H2O2 concentrations. Solid lines: model predictions.
[H2O2]0 5 mM. [4 MC]0 0.1 mM.

Appendix A. Supplementary data

Supplementary data related to this article can be found at http://


dx.doi.org/10.1016/j.chemosphere.2016.06.110.

References

Ahel, M., Schaffner, C., Giger, W., 1996. Behaviour of alkylphenol polyethoxylate
surfactants in the aquatic environment - III. Occurrence and elimination of their
persistent metabolites during inltration of river water to groundwater. Water
Res. 30, 37e46.
Benjamin, M.M., 2002. Water Chemistry. Waleland Press Inc., p. 139
Bielski, B.H.J., Cabelli, D.E., Arudi, R.L., Ross, A.B., 1985. Reactivity of HO2/O2 radicals
in aqueous solution. J. Phys. Chem. Ref. Data 14, 1041e1100.
Boltz, D., Holwell, J., 1978. Colorimetric Determination of Nonmetals, second ed.
John Wiley and Sons, New York.
Boonrattanakij, N., Lu, M., Anotai, J., 2009. Kinetics and mechanism of 2,6-dimethy-
aniline degradation by hydroxyl radicals. J. Hazard. Mater. 172, 952e957.
Buxton, G.V., Greenstock, C.L., Helman, W.P., Ross, A.B., 1988. Critical review of rate
constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl
radicals (OH/O) in aqueous solution. J. Phys. Chem. Ref. Data 17, 513e886.
Crittenden, J.C., Hu, S., Hand, D.W., Green, S.A., 1999. A kinetic model for H2O2/UV
Fig. 10. Evolution of TOC with time. Solid line: model prediction. [PC]0 0.5 mM, process in a completely mixed batch reactor. Wat. Res. 33, 2315e2328.
Davies, M.S., Mile, B., Rowlands, C.C., 1995. Oxidation of 4-methylcatechol by
[H2O2]0 5 mM.
dioxygen studies by ESR spectroscopy. The different regioselectivity of OH and
MeO nucleophilic attack and kinetic deuterium isotope effects. Magn. Reson.
Chem. 33, 15e19.
to over-estimate degradation rates at higher H2O2 concentrations. De Luis, A., Lombran ~ a, J.I., Menendez, A., 2011. Modeling of the radicalary state in
the H2O2/UV oxidation system to predict the degradation kinetics of phenolic
mixture solutions. AIChE J. 30, 196e207.
Dong, B., Kahl, A., Cheng, L., Vo, H., Ruehl, S., Zhang, T., Snyder, S., S aez, A.E.,
4. Concluding remarks Quanrud, D., Arnold, R.G., 2015. Fate of trace organics in a wastewater efuent
dependent stream. Sci. Total Environ. 518e519, 479e490.
Eriksen, T.E., Lind, J., Merenyi, G., 1985. On the acid-base equilibrium of the car-
A comprehensive kinetic model was developed to quantify the bonate radical. Radiat. Phys. Chem. 26, 197e199.
oxidation of p-cresol by UV/H2O2 AOP in a batch reactor. The model Ershov, B.G., Janata, E., Alam, M.S., Gordeev, A.V., 2008. Studies of the reaction of the
was based on known elementary chemical and photo-chemical hydroxyl radical with the oxalate ion in an acidic aqueous solution by pulse
radiolysis. Russ. Chem. Bull. Int. Ed. 57, 1189.
reactions of the UV/H2O2 photolysis system, together with experi- Getoff, N., Schworer, F., Markovie, V.M., Sehested, K., Nielsen, S.O., 1971. Pulse
mentally established direct photolysis reactions and proposed OH- radiolysis of oxalic acid and oxalates. J. Phys. Chem. 75, 749e755.
radical reactions with intermediates. The kinetics of p-cresol Glaze, W.H., Lay, Y., Kang, J.-W., 1995. Advanced oxidation processes. A kinetic
model for the oxidation of 1,2-dibromo-3-chloropropane in water by the
decomposition, as well as the main reaction intermediate (4- combination of hydrogen peroxide and UV radiation. Ind. Eng. Chem. Res. 34,
methylcatechol) was successfully simulated throughout a wide 2314e2323.
range of operating conditions. The approach employed shows that Greenwood, N.N., Earnshaw, A., 1997. Chemistry of the Elements, second ed. But-
terworth-Heinemann, p. 310.
the full reaction mechanism needs to be considered to evaluate
Gohn, M., Getoff, N., 1977. Pulse radiolysis of 3,4-dihydroxytoluene. J. Chem. Soc. 73,
process performance. In particular, reaction intermediates, espe- 1207e1215.
cially those formed before ring cleavage, play an important role in Hart, E.J., 1951. Mechanism of the g-Ray induced oxidation of formic acid in aqueous
the quenching of hydroxyl radicals and, consequently, those com- solution. J. Am. Chem. Soc. 73, 68e73.
Holeman, J., Bjergbakke, E., Sehested, K., 1987. The importance of radical-radical
pounds must be considered in kinetic models of AOP degradation of reactions in pulse radiolysis of aqueous carbonate/bicarbonate. Proc. Tihany
phenolic compounds. Symp. Radiat. Chem. 1, 149e153.
T. Zhang et al. / Chemosphere 161 (2016) 349e357 357

Housecroft, C.E., Sharpe, A.G., 2005. Inorganic Chemistry, second ed. Prentice- Tetrahedron 24, 3475e3487.
Pearson-Hall, p. 368. Perry, R., Green, D., Maloney, J., 1981. Perrys Chemical Engineerings Handbook,
Huang, D., Zhang, X., Chen, Z.M., Zhao, Y., Shen, X.L., 2011. The kinetics and fth ed. McGraw-Hill, New York.
mechanism of an aqueous phase isoprene reaction with hydroxyl radical. Pospisil, J., Ettel, V., Skola, V., 1957. Pyrocatechol oxidation to muconic acid. Chem.
Atoms. Chem. Phys. 11, 7399e7415. Prumysl. 7, 244e248.
Jin, F., Zhong, H., Cao, J., Cao, J., Kawasaki, K., Kishita, A., Matsumoto, T., Tohji, K., Rojas, M.R., Perez, F., Whitley, D., Arnold, R.G., S aea, A.E., 2010. Modeling of
Enomoto, H., 2010. Oxidation of unsaturated carboxylic acids under hydro- advanced oxidation of trace organic contaminants by hydrogen peroxide
thermal conditions. Bioresour. Technol. 101, 7624e7634. photolysis and Fentons reaction. Ind. Eng. Chem. Res. 49, 11331e11343.
Jongberg, S., Gislason, N.E., Lund, M.N., Skibsted, L.H., Waterhouse, A.L., 2011. Thiol- Rojas, M.R., Leung, C., Whitley, D., Zhu, Y., Arnold, R.G., Sa ea, A.E., 2011. Advanced
quinone adduct formation in myobrillar proteins detected by LC-MS. J. Agric. oxidation of trace organics in water by hydrogen peroxide solar photolysis. Ind.
Food Chem. 59, 6900e6905. Eng. Chem. Res. 50, 12479e12487.
Joschek, H.I., Miller, S.I., 1966. Photooxidation of phenol, cresols, and dihydrox- Scheck, C.K., Frimmel, F.H., 1995. Degradation of phenol and salicylic acid by ul-
ybenzenes. J. Am. Chem. Soc. 88 (14), 3273e3281. traviolet radiation/hydrogen peroxide/oxygen. Wat. Res. 29, 2346e2352.
Kang, N., Lee, D.S., Yoon, J., 2002. Kinetic modeling of Fenton oxidation of phenol Scullion, S.D., Clench, M.R., Cooke, M., Ashcroft, A.E., 1996. Determination of sur-
and monochlorophenols. Chemosphere 47, 915e924. factants in surface water by solid-phase extraction, liquid chromatography and
Lay, Y.S., 1989. Oxidation of 1,2-dibromo-3-chloropropane in Groundwater Using liquid chromatography-mass spectrometry. J. Chromatogr. A 733, 207e216.
Advanced Oxidation Processes. Ph.D. dissertation. University of California, Los Sehested, K., Cortzen, H., Christensen, H.C., Hart, E.J., 1975. Rates of reaction of O,
Angeles. OH and H with methylated benzenes in aqueous solution. Optical spectra of
Lee, P.-C., Lee, W., 1996. In vivo estrogenic action of nonylphenol in immature fe- radicals. J. Phys. Chem. 79, 310e315.
male rates. Bull. Environ. Contam. Toxicol. 57, 341e348. Soto, A.M., Justicia, H., Wray, J.W., Sonnenschein, C., 1991. p-Nonyl-phenol: an es-
Leitner, N.K.V., Dore, M., 1997. Mechanism of the reaction between hydroxyl radicals trogenic xenobiotic released from modied polystyrene. Environ. Health
and glycolic, glyoxylic, acetic and oxalic acids in aqueous solution: consequence Perspect. 92, 167e173.
on hydrogen peroxide consumption in the H2O2/UV and O3/H2O2 systems. Stefan, M.I., Hoy, A.R., Bolton, J.R., 1996. Kinetics and mechanism of the degradation
Water Res. 31, 1383e1397. and mineralization of acetone in dilute aqueous solution sensitized by the UV
Liao, C.-H., Gurol, M., 1995. Chemical oxidation by photolytic decomposition of photolysis of hydrogen peroxide. Environ. Sci. Technol. 30, 2382e2390.
hydrogen peroxide. Environ. Sci. Technol. 29, 3007e3014. Song, W., Ravindran, V., Pirbazari, M., 2008. Process optimization using a kinetic
Loyo-Rosales, J.E., Rice, C.P., Torrents, A., 2007. Fate of octyl- and nonylphenol model for the ultraviolet radiation-hydrogen peroxide decomposition of natural
ethoxylates and some carboxylated derivatives in three American wastewater and synthetic organic compounds in groundwater. Chem. Eng. Sci. 63,
treatment plants. Environ. Sci. Technol. 41, 6815e6821. 3249e3270.
Lye, C.M., Frid, C.L.J., Gill, M.E., Cooper, D.W., Jones, D.M., 1999. Estrogenic alkyl- Sun, W., Saeys, M., 2008. First principles study of the reaction of formic and acetic
phenols in sh tissues, sediments, and waters from the UK. Tyne Tees Estuaries. acids with hydroxyl radicals. J. Phys. Chem. A 112, 6918e6928.
33, 1009e1014. Swartz, C.H., Reddy, S., Benotti, M.J., Yin, H., Barber, L.B., Brownawell, B.J., Rudel, R.A.,
MacDonald, B.C., Lvin, S.J., Patterson, H., 1997. Correction of uorescence inner lter 2006. Steroid estrogens, nonylphenol ethoxylate metabolites, and other
effects and the partitioning of pyrene to dissolved organic carbon. Anal. Chim. wastewater contaminants in groundwater affected by a residential septic sys-
Acta 338, 155e162. tem on Cape Cod, MA. Environ. Sci. Technol. 40, 4894e4902.
Matsuura, T., Omura, K., 1974. Photochemical hydroxylation of aromatic com- Volman, D.H., Chen, J.C., 1959. The photochemical decomposition of hydrogen
pounds. Synthesis 3, 173e184. peroxide in aqueous solutions of allyl alcohol at 2537 . J. Am. Chem. Soc. 81
Moreira, J., Serrano, B., Ortiz, A., de Lasa, H., 2012. A unied kinetic model for phenol (16), 4141e4144.
photocatalytic degradation over TiO2 photocatalysts. Chem. Eng. Sci. 78, Waite, J.H., 1976. Calculating extinction coefcients for enzymatically produced 0-
186e203. quinones. Anal. Biochem. 75, 211e218.
Murcia, M.D., Vershinin, N.O., Briantceva, N., Gomez, M., Gomez, E., Cascales, E., Wang, K., Hsieh, Y., Chen, L., 1998. The heterogeneous photocatalytic degradation,
Hidalgo, A.M., 2015. Development of a kinetic model for the UV/H2O2 photo- intermediates and mineralization for the aqueous solution of cresols and
degradation of 2,4-dichlorophenoxiacetic acid. Chem. Eng. J. 266, 356e367. nitrophenols. J. Hazard. Mater. 59, 251e260.
Nagarnaik, P.M., Boulanger, B., 2011. Advanced oxidation of alkylphenol ethoxylates Waters, W.A., 1964. Mechanism of Oxidation of Organic Compounds. Methuen,
in aqueous systems. Chemosphere 85, 854e860. p. 132.
Neta, P., Huie, R.E., Ross, A.B., 1988. Rate constants for reactions of inorganic radicals Weinstein, J., Bielski, B.H.J., 1979. Kinetics of the interaction of perhydroxyl and
in aqueous solution. J. Phys. Chem. Ref. Data. 17, 1027e1284. superoxide radicals with hydrogen peroxide. The Haber-Weiss reaction. J. Am.
Olariu, R.I., Klotz, B., Barnes, I., Becker, K.H., Mocanu, R., 2002. FT-IR study of the Chem. Soc. 101, 58e62.
ring-retaining products from the reaction of OH radicals with phenol, o-, m-, Wojna rovits, L., Foldiak, G., Dangelantonio, M., Emmi, S.S., 2002. Mechanism of OH
and p-cresol. Atmos. Environ. 36, 3685e3697. radical-induced oxidation of p-cresol to p-methylphenolxyl radical. Res. Chem.
Omura, K., Matsuura, T., 1968. Photo-induced reactions-IX: the hydroxylation of Intermed. 28, 373e386.
phenols by the photo-decomposition of hydrogen peroxide in aqueous media.

Das könnte Ihnen auch gefallen